You are on page 1of 8

JOURNAL OF NANOSCIENCE AND NANOTECHNOLOGY

Biodegradable Polylactide/Montmorillonite
Nanocomposites
Suprakas Sinha Ray,a, * Kazunobu Yamada,b Masami Okamoto,a, * and Kazue Uedab
a
Advanced Polymeric Materials Engineering, Graduate School of Engineering, Toyota Technological Institute, Nagoya 468 851, Japan
b
Unitika Ltd., Kyoto 611 0021, Japan

Our continuing research on the preparation, characterization, materials properties, and biodegrad-
ability of polylactide (PLA)/organically modified layered silicate (OMLS) nanocomposites has yielded
results on PLA /montmorillonite nanocomposites. Montmorillonite (mmt) modified with dimethyldioc-
tadecylammonium cation was used as an OMLS for nanocomposite preparation. The internal struc-
ture of nanocomposites on the nanometer scale was established with the use of wide-angle X-ray
diffraction patterns and transmission electron micrographic observation. All nanocomposites exhib-
ited significant improvement in crystallization behavior, mechanical properties, flexural properties,
heat distortion temperature, and O2 gas permeability when compared with pure PLA.

Keywords: Polylactide, Montmorillonite, Nanocomposite, Crystallization, Materials Properties.

1. INTRODUCTION in mequiv/100 g). This charge is not locally constant, as it


varies from layer to layer and must rather be considered as
Over the last few years, polymer/layered silicate nanocom- an average value over the whole crystal. Layered silicates
posites have attracted great interest from researchers, both in have two types of structure, tetrahedral-substituted and octa-
industry and in academia, because they often exhibit remark- hedral-substituted. In case of tetrahedral-substituted layered
able improvement of materials properties when compared silicates the negative charge is located on the surface of sili-

RESEARCH ARTICLE
with pure polymer or conventional composites (both micro- cate layers, and, hence, the polymer matrices can react with
and macro-composites). These improvements include higher tetrahedral-substituted silicate more compared with that of
moduli,1 increased strength and heat resistance,2 decreased octahedral-substituted.7
gas permeability3 and flammability,4 and increased biodegra- Pristine layered silicates usually contain hydrated Na
dability of biodegradable polymers.5 On the other hand, or K ions.8 Obviously, in this pristine state layered sili-
these materials have also been proved to be unique model cates are only miscible with hydrophilic polymers, such as
systems for the study of the structure and dynamics of poly- poly(ethylene oxide) (PEO),9 poly(vinyl alcohol) (PVA),10
mers in confined environments.6 etc. To render layered silicates miscible with other polymer
The commonly used layered silicates for the preparation matrices, one must convert the normally hydrophilic silicate
of polymer/layered silicate nanocomposites belong to the surface to organophilic, which makes possible the interca-
same general family of 2:1 layered or phyllosilicates.7 Their lation of many engineering polymers. Generally, this can be
crystal structure consists of layers made up of two silica done by ion-exchange reactions with cationic surfactants,
tetrahedra fused to an edge-shared octahedral sheet of either including primary, secondary, tertiary, and quaternary alkyl-
aluminum or magnesium hydroxide. The layer thickness is ammonium or alkylphosphonium cations. The role of alkyl-
around 1 nm, and the lateral dimensions of these layers may ammonium or alkylphosphonium cations in the organosil-
vary from 30 nm to several micrometers and even larger, icates is to lower the surface energy of the inorganic host
depending on the particular layered silicate. Stacking of the and to improve the wetting characteristics with the polymer
layers leads to a regular van der Waals gap between the matrix; this results in a larger interlayer spacing. Addi-
layers called the interlayer or gallery. Isomorphic substitu- tionally, the alkylammonium or alkylphosphonium cations
tion within the layers (for example, Al3 replaced by Mg2 could provide functional groups that can react with the
or by Fe2, or Mg2 replaced by Li) generates negative polymer matrix or in some cases initiate the polymeriza-
charges that are counterbalanced by alkali and alkaline earth tion of monomers to improve the strength of the interface
cations situated inside the galleries. The type of layered sili- between the inorganic and the polymer matrix.11, 12
cate is characterized by a moderate surface charge (known The main reason for the improved properties of pure poly-
as cation exchange capacity (CEC), and generally expressed mer in nanocomposites is interfacial interaction between
polymer matrix and organically modified layered silicate
*Authors to whom correspondence should be addressed. (OMLS) as opposed to conventional composites. A few
J. Nanosci. Nanotech. 2003, Vol. 3, No. 6 © 2003 by American Scientific Publishers 1533-4880/2003/03/001/008/$17.00+.25 doi:10.1166/jnn.2003.220 1
Ray et al./Biodegradable Polylactide/Montmorillonite J. Nanosci. Nanotech. 2003, 3, 1–5

weight percent of OMLS that is properly dispersed through- ically modified montmorillonite (C2C218-mmt) used in this
out the matrix thus creates a much higher surface area for study was supplied by Hojun Yoka Co., Japan, and was
polymer-filler interfacial interactions than do conventional synthesized by replacing Na in montmorillonite (mmt)
composites.13 On the basis of the strength of polymer- (original thickness of 1 nm and average length of 100 nm)
OMLS interactions, three different types of nanocomposites of a CEC of 90 mequiv/100 g with dimethyldioctadecy-
are structurally achievable: intercalated nanocomposites, lammonium cation by ion exchange reaction.
where insertion of polymer chains into the layered silicate
structure occurs in a crystallographically regular fashion, 2.2. Nanocomposite Preparation
regardless of the polymer-to-OMLS ratio, and a repeat dis-
tance of few nanometers; flocculated nanocomposites, For nanocomposite preparation, C2C218-mmt (powder form)
where intercalated stacked silicate layers eventually floc- and PLA (pellet form) were first dry-mixed by shaking
culate because of the hydroxylated edge-edge interactions; them in a bag. The mixture was then melt-extruded with a
and exfoliated nanocomposites, where the individual sili- twin-screw extruder (PCM-30, Ikegai machinery Co.) oper-
cate layers are separated in a polymer matrix by average ated at 210 °C30 (screw speed  100 rpm, feed rate  120
distances that totally depend on the clay loading.14 g/min) to yield nanocomposite strands. Henceforth, the
Recently, the development of biodegradable polymeric product nanocomposites are abbreviated as PLACNs.
materials with excellent materials properties is a subject of PLACNs prepared with three different amounts of C2C218-
great research challenge in materials science. One of the mmt of 4, 5, and 7 wt% are correspondingly abbreviated
most promising polymers in this direction is polylactide as PLACN4, PLACN5, and PLACN7, respectively. The
(PLA) because it is made from renewable agriculture strands were pelletized and dried under vacuum at 60 °C for
products and is readily biodegradable.15, 16 PLA is linear 48 h to remove water.
aliphatic thermoplastic polyester, produced by the ring- The dried PLACN pellets were then converted into sheets
opening polymerization of lactide.17–19 Lactide is a cyclic with a thickness of 0.7–2 mm by pressing with 1.5 MPa at
dimer prepared by the controlled depolymerization of lac- 190 °C for 3 min. The molded sheets were then quickly
tic acid, which in turn is obtained by the fermentation of quenched between glass plates and then annealed at 110 °C
corn, sugar beet, etc.15, 16 It has very good mechanical for 1.5 h to crystallize isothermally before being subjected to
properties, thermal plasticity, fabric ability, and biocom- wide-angle X-ray diffraction (WAXD), transmission elec-
patibility.20–21 Even when burned it produces no nitrogen tron microscopy (TEM), and dynamic mechanical property
oxide gases and only one-third the combustible heat gener- measurements.
RESEARCH ARTICLE

ated by polyolefin, and does not damage incinerators and,


thus, provides significant energy savings.22 So PLA is a 2.3. Characterization
promising polymer for various end-use applications, and
currently there is increasing interest in using PLA for dis- 2.3.1. WAXD
posable degradable plastic articles.23 However, some other WAXD analyses were performed for the C2C218-mmt
properties such as heat distortion temperature (HDT), flex- powder and three different PLACN sheets with an MXlabo
ural and gas barrier properties, and melt viscosity for fur- X-ray diffractometer (MAC Science Co., generator of 3 kW,
ther processing are frequently not good enough for wide- a graphite monochromator, CuKa radiation (wavelength,
ranging applications.24 On the other hand, preparation of l  0.154 nm), operated at 40 kV/20 mA). The samples
naocomposites with OMLS has already proved to be an were scanned in fixed-time (FT) mode with a counting
effective way to improve these properties significantly.1–6 time of 2 s under a diffraction angle of 2v in the range of
In our previous papers,25–29 we have reported on the 1° to 70°.
preparation, characterization, and materials properties of
various kinds of PLA/OMLS nanocomposites. In this
2.3.2. TEM
paper we describe the preparation, characterization, crys-
tallization behavior, mechanical and material properties of To clarify the nanoscale structure of various PLACNs,
another type of PLA nanocomposite prepared with dimeth- high-resolution TEM (H-7100; Hitachi Co.) was also used,
yldioctadecylammonium-modified montmorillonite with a at an accelerating voltage of 100 kV. An ultrathin section
simple melt extrusion technique. of crystallized sheet (perpendicular to the compression
mold) with a thickness of 100 nm was microtomed at
80 °C with a Reichert Ultra cut cryo-ultramicrotome
2. EXPERIMENTAL DETAILS
without staining.
2.1. Materials
2.3.3. Gel Permeation Chromatography
PLA with D content of 1.1–1.7% (supplied by Unitika Co.
Ltd., Japan) was dried under vacuum at 60 °C and kept Weight-averaged (Mw) and number-averaged (Mn) molec-
under dry nitrogen gas for 1 week prior to use. The organ- ular weights of PLA (before and after nanocomposite
2
J. Nanosci. Nanotech. 2003, 3, 1–5 Ray et al./Biodegradable Polylactide/Montmorillonite

preparation) were determined by gel permeation chroma- and placed on a laboratory hot plate at 200 °C for 60 s
tography (GPC) (LC-VP, Shimadzu Co.), with polystyrene to obtain a thin film 30 mm in thickness. The molten film
standards for calibration and tetrahydrofuran (THF) as the was then quickly quenched to the desired temperature (100
carrier solvent at 40 °C, with a flow rate of 0.5 ml/min. For °C) by putting it on a thermostatted hot stage (Linkam
the GPC measurements first PLA or nanocomposites were RTVMS; Linkam Scientific Instruments, Ltd.) mounted on
dissolved in CHCl3 and then diluted with THF. In the case a POM (Nikon OPTIPHOTO2-POL). After complete crys-
of nanocomposites, after dissolving in CHCl3 the insoluble tallization, the nature of crystallite growth was observed
clay particles were separated from the dissolved PLA by with a POM.
filtration.
2.3.6. Dynamic Mechanical Analysis
2.3.4. Differential Scanning Calorimetry
Dynamic mechanical properties of the pure PLA and
The glass transition (Tg), melting (Tm), and crystallization PLACNs were measured with a Reometrics Dynamic Ana-
(Tc) temperatures, as well as degree of crystallinity (xc) lyzer (RDAII) in the tension-torsion mode. The tempera-
of pure PLA and PLACNs, were determined by tempera- ture dependence of dynamic storage modulus (G), loss
ture-modulated differential scanning calorimetry (DSC) modulus (G), and their ratio, tan d, of pure PLA and vari-
(TMDSC) (MDSC, TA2920, TA instruments), operated at ous PLACNs were determined at a constant frequency (q)
a heating rate of 5 °C/min with a heating/cooling cycle of of 6.28 rad/s with a strain amplitude of 0.05%, and in the
the modulation period of 60 s and an amplitude of 0.769 temperature range of 20 to 160 °C with a heating rate of
°C. For the measurement of xc prior to DSC analysis, the 2 °C/min.
extra heat absorbed by the crystallites formed during heat-
ing had to be subtracted from the total endothermic heat
2.3.7. Flexural Properties and Heat Distortion
flow due to the melting of the whole crystallites. This can
Temperature
be done according to the principles and procedures
described in our previous paper.31 By considering the The sample pellets were injection-molded with an injec-
melting enthalpy of 100% crystalline poly(L-lactide) as tion machine (IS-80G; Toshiba Machinery Co.) operated
93 J/g,32 we have estimated the value of the xc for pure at 190 °C with a mold temperature of 30 °C. Flexural mod-
PLA and PLACNs, which is presented in Table I. ulus and strength of the injection-molded specimens (thick-
ness 3.2 mm, annealed at 120 °C for 30 min) were mea-

RESEARCH ARTICLE
2.3.5. Light Scattering and Polar Optical Micrographic sured according to the ASTM D-790 method (model 2020;
Observations Intesco Co.) with a strain rate of 2 mm/min at room tem-
perature (25 °C). The heat distortion temperature (HDT)
To investigate the crystallite texture, the pure PLA and tests of neat PLA and various PLACNs were conducted
PLACN samples were subjected to light scattering (LS) with crystallized injection molded samples (HDT Tester;
experiments under Hv scattering mode with the radiation Toyoseiki Co.) according to the ASTM D-648 method
of a polarized He-Ne laser at 632.8-nm wavelength. The with a heating rate of 2 °C/min. We also conducted load
details of the LS measurement were described in our previ- dependence of HDT of neat PLA and PLACN7 by the
ous paper.33 same procedure.
We also observed the crystallite growth behavior of pure
PLA and PLACNs with a polar optical micrograph (POM).
2.3.8. Measurement of O2 Gas Transmission Rates
Dried pellets were sandwiched between two glass slides
Oxygen gas transmission rates of pure PLA and various
PLACNs were measured at 20 °C and 90% relative humid-
Table I. Characteristic parameters of neat PLA and various PLACNs.
ity by the ASTM D-1434 differential pressure method
Parameters Neat PLA PLACN4 PLACN5 PLACN7 (GTR-30XAU; Yanaco Co.). Test samples were prepared
3
Mw  10 (gmol ) 1
177 a
163 158 155
by compression molding (thickness 300 mm), and melt-
Mw/Mn 1.58 1.61 1.60 1.59 quenched samples were used for this measurement.
Tg (°C) 60 58.6 57.4 57.3
Tm (°C) 168 170 169.6 169.6
Tc (°C) 127.2 101 101.7 100.6 3. RESULTS AND DISCUSSION
xc (%) 36 47.5 47.2 51.7
N  105 (mm3) 2.7 822.6 337.4 207.9 3.1. Nanocomposite Structure
d001 (nm) — 3 2.95 2.85
D (nm)b — 11.3 14.5 15.9 The structure of the nanocomposites in the nanometer
D/d001 — 4 5 6 range has typically been elucidated by WAXD and TEM.
a
Extruded PLA under the same condition. WAXD provides direct evidence of the intercalation of
b
Calculated by using Scherrer equation. the polymer chains into the silicate galleries. On the other
3
Ray et al./Biodegradable Polylactide/Montmorillonite J. Nanosci. Nanotech. 2003, 3, 1–5

hand, TEM offers a qualitative understanding of the inter- gradually increases with increasing clay loading. There-
nal structure through direct visualization. WAXD patterns fore, we can conclude that the well-ordered intercalated
for the pure C2C218-mmt powder and three different PLACNs were formed and coherence order of the silicate
PLACNs are presented in Figure 1. The interlayer gallery layers increases with increasing C2C218-mmt content. Divid-
height, calculated as the difference of the d001 distances ing the value of D by the d001 value of each PLACN, we
obtained by WAXD and the individual layer thickness for can estimate the number of the stacked individual silicate
the C2C218-mmt, is 2.5 nm, which, upon intercalation by layers to be about 4 for PLACN4, 5 for PLACN5, and
PLA, expands to 3 nm in the case of PLACN4. With about 6 in the case of PLACN7.
increasing clay content this distance gradually decreases, On the other hand, TEM is a qualitative tool used for the
and the value of d001 is equal to 2.85 nm in the case of determination of overall clay dispersion. Figure 2 shows
PLACN7. From the WAXD patterns, the crystallite size the typical bright-field TEM image of PLACN5 in which
(D) of intercalated stacked silicate layers of each PLACN dark entities are the cross section of the stacked and inter-
is calculated with the Scherrer equation.34 The calculated calated silicate layers.35 From the TEM image, we observed
value of D for each PLACN is also presented in Table I. It the large anisotropy of the stacked and flocculated silicate
is clearly established that the crystallite size (the thickness layers, which have an original thickness of 1 nm and an
of the dispersed stacked silicate layers) in the PLACNs average length of 100 nm.

3.2. GPC

GPC results of PLA in the pure state or in C2C218-mmt-


filled systems are also presented in Table I. As anticipated,
the incorporation of C2C218-mmt resulted in a reduction in
the molecular weight of the PLA matrix. Decreased mole-
cular weights of PLA in nanocomposites may be explained
by the shear mixing of PLA and C2C218-mmt, resulting
in a certain extent of hydrolysis of PLA matrix at high
temperature.
RESEARCH ARTICLE

3.3. Thermal Properties and Crystallite Morphology

Figure 3 shows DSC traces for the melt-quenched samples


of neat PLA and three different PLACNs. These data are
obtained from the first run and during the heating process.
Samples were prepared by using a hot press. Neat PLA
and PLACN pellets were melted at 190 °C hold for 3 min

Fig. 1. WAXD patterns of C2C218-mmt powder and various PLACNs.


The dashed line indicates the location of the silicate (001) reflection of Fig. 2. High-resolution TEM bright-field image of crystallized PLACN5
C2C218-mmt. The asterisks indicate the (001) peak for C2C218-mmt dis- sheet. The dark entities are the cross section of the intercalated mmt layers,
persed in PLA matrix. and the bright areas are the matrix.

4
J. Nanosci. Nanotech. 2003, 3, 1–5 Ray et al./Biodegradable Polylactide/Montmorillonite

Fig. 3. DSC scans of the melt-quenched samples of pure PLA and vari-
ous PLACNs.

at the same temperature under 1.5 MPa pressure, and


then quickly quenched between glass plates. From Figure
3 we can see an endothermic peak for all samples in the
temperature range of 55–60 °C. The temperature corre-
sponding to the endothermic peak for each sample is con-
sidered to be Tg of PLA. For all samples at Tg there is a
steplike change, which is due to enthalpy relaxation. This
type of steplike change at Tg has also been recognized for
some other polymers.36

RESEARCH ARTICLE
On the other hand, all samples show an exothermic
peak, and that can be correlated to the crystallization of
PLA in every sample; the corresponding temperature is
known as the crystallization temperature, Tc. In the case of Fig. 4. (a) Polarized optical micrographs of pure PLA and various
PLACNs this peak is sharper and appeared at lower tem- PLACNs isothermally crystallized at 100 °C for 1.5 h. (b) Hv-light scat-
perature than that of neat PLA (see Table I). Therefore, tering patterns for pure PLA and various PLACNs isothermally crystal-
lized at 110 °C for 1.5 h.
C2C218-mmt seems to enhance the rate of crystallization of
PLA. It should be noted here that Tc does not depend upon
C218C2-mmt. This behavior suggests that a small amount
of C2C218-mmt is enough to serve as a nucleating agent for less organized crystallites. From LS patterns, the number
crystallization. A systematic DSC study of all samples of heterogeneous nuclei N can be estimated from a rough
shows that the Tg signal of neat PLA gradually weakens approximation (i.e., all of the crystallites are of identical
with increasing C2C218-mmt content. This behavior sug- size). The primary average nucleation density of the crys-
gests that C2C218-mmt layers affect all polymers, and there tallites, N, is given by37
is very small amount of bulk-like PLA present to manifest N  (3/4p)(Dm/2)3
itself through the thermal transition.
POM photographs of neat PLA and various PLACNs where Dm is the average maximum diameter of the crystal-
are presented in Figure 4a. All samples were crystallized at lite, that is, the attainable diameter before impingement.
100 °C beforehand. Neat PLA exhibits well-defined large The calculated values of N at 100 °C are also presented in
crystallite morphology, whereas the size of the PLACN Table I. This behavior indicates that the surface of the dis-
crystallites is significantly smaller. This observation indi- persed C2C218-mmt layers acts as a nucleating agent for
cates that the surface of dispersed C2C218-mmt has a very PLA crystallization, which is evident from the increase in
strong effect on PLA crystallization. the number of density of nuclei causing smaller crystallite
This behavior is more clearly seen in LS patterns (see formation.37 So our investigation explores the role of
Fig. 4b), where, for PLACNs, a clear large smeared four- C2C218-mmt as a nucleating agent for PLA crystallization.
leaf-clover pattern is observed compared with the crystal- Such a discussion is beyond the objective of this paper, and
lized neat PLA, indicating formation of a large number of we will report it separately.38
5
Ray et al./Biodegradable Polylactide/Montmorillonite J. Nanosci. Nanotech. 2003, 3, 1–5

3.4. DMA prominent and hence a strong enhancement of the modulus


appears.14, 39
DMA measures the response of a given material to an oscil- On the other hand, above Tg the enhancement of G is
latory deformation (here in tension-torsion mode) as a func- significant in the intercalated PLACNs in comparison with
tion of the temperature. DMA results are expressed by three that below Tg, indicating a plastic response to the deforma-
main parameters: (a) the storage modulus (G), correspond- tion is prominent in the presence of clay when material
ing to the elastic response to the deformation; (b) the loss becomes soft. However, the presence of clay particles does
modulus (G), corresponding to the plastic response to the not lead to a significant shift and broadening of the tan d
deformation, and (c) tan d, that is, the G/G ratio, which is curves for all PLACNs compared with that of pure PLA.
useful for determining the occurrence of molecular mobil- This behavior has been ascribed to the restricted segmental
ity transitions such as the glass transition temperature. motions at the organic–inorganic interface neighborhood
Here DMA analysis has been studied to track the tem- of intercalated PLACNs.
perature dependence of storage modulus upon PLACN
formation. Figure 5 shows the temperature dependence of
G, G, and tan d for PLA and corresponding PLACNs. For 3.5. Flexural Properties and HDT
all PLACNs, significant enhancement of G can be seen in In Table II, we report the flexural modulus, flexural strength,
the investigated temperature range, indicating that interca- and distortion at break of neat PLA and three PLACNs mea-
lated PLACN is strongly influenced by the elastic proper- sured at 25 °C. There is a significant increase of flexural
ties of the PLA matrix. modulus for PLACN4 compared with that of pure PLA fol-
Below Tg, the enhancement of G is clear in the inter- lowed by a constant value with increasing C2C218-mmt con-
calated PLACNs, but at the temperature range of 80 °C tent. On the other hand, flexural strength and distortion at
to 145 °C all three nanocomposites exhibit much higher break exhibited a remarkable increase with PLACN4 and
enhancement of G as compared with that of pure PLA. then gradually decreased with C2C218-mmt loading. This
This is due to the mechanistic reinforcement by clay parti- behavior may be due to the high C2C218-mmt content, which
cles at high temperature. Above Tg, when materials become leads to a brittleness of materials. Therefore, we can control
soft, the reinforcement effect of the clay particles becomes the flexural strength and distortion by increasing or decreas-
ing OMLS content, and incorporation of C2C218-mmt of
around 4 wt% is the optimum for achieving a high value of
flexural strength and distortion.
RESEARCH ARTICLE

The nanodispersion of C2C218-mmt in pure PLA also pro-


motes a higher HDT. We examined the HDT of pure PLA
and various PLACNs with different load conditions. As
seen in Figure 6a, in the case of PLACN4, there is a marked
increase in HDT with an intermediate load of 0.98 MPa,
from 76 °C for the pure PLA to 91.3 °C for PLACN4. The
value of HDT gradually increases with increasing clay
content, and, in the case of PLACN7, the value increases to
107 °C.
On the other hand, imposed load dependence on HDT is
clearly observed in the case of PLACNs. Figure 6b shows
the typical load dependence in the case of PLACN7. The
increase in HDT of neat PLA due to nanocomposite prepara-
tion is a very important property improvement, not only from
the industrial point of view but also for molecular control on
the silicate layers, that is, crystallization through interaction
between PLA molecules and SiO4 tetrahedral layers.
In the case of high load (1.81 MPa), it is very difficult to
achieve high HDT enhancement without strong interaction

Table II. Flexural properties of neat PLA and various PLACNs.

Flexural properties Neat PLA PLACN4 PLACN5 PLACN7

Modulus (GPa) 4.84 5.43 5.38 5.39


Strength (MPa) 86 102 107 99
Fig. 5. Temperature dependence of storage modulus (G), loss modulus Distortion (%) 1.9 3.9 2.1 1.9
(G), and their ratio, tan d, for pure PLA and various PLACNs.

6
J. Nanosci. Nanotech. 2003, 3, 1–5 Ray et al./Biodegradable Polylactide/Montmorillonite

Fig. 7. WAXD patterns of purePLA and various PLACNs. Samples were


crystallized at 110 °C for 1.5 h before measurements. The data were verti-
cally offset to make a clear presentation.

The above expression assumes that the sheets are placed


such that the sheet plane is perpendicular to the diffusive
Fig. 6. (a) Organoclay (C2C218-mmt) (wt%) dependence of HDT of pure pathway. The O2 gas permeability for the pure PLA and
PLA and various PLACNs. (b) Load dependence of HDT of pure PLA and various PLACNs are presented in Figure 8. As seen from
PLACN7.
Figure 8, the permeability of O2 gas through the PLACN
films improved compared with pure PLA film.

between polymer matrix and OMLS.40 In the case of all


PLACNs studied here, the values of Tm (cf. Table I) do not 4. CONCLUSIONS

RESEARCH ARTICLE
change significantly as compared with that of pure PLA. We have successfully prepared biodegradable polylactide/
Furthermore, in WAXD analyses up to 2v  70° (see Fig. 7), C2C218-mmt nanocomposites by simple melt extrusion of
we observed no big shift or formation of new peaks in the PLA and C2C218-mmt, wherein silicates of C2C218-mmt
crystallized PLACNs. So the improvement of HDT with are intercalated, stacked and flocculated, and nicely dis-
intermediate load (0.98 MPa) originates from the better tributed in the PLA matrix. All nanocomposites exhibited
mechanical stability of the PLACNs due to mechanical significant improvement in mechanical and various other
reinforcement by the dispersed clay particles, higher value materials properties when compared with those of neat
of xc, and intercalation. This is qualitatively different from PLA. These improvements include rate of crystallization,
the behavior of nylon-6/OMLS nanocomposites, where the mechanical and flexural properties, heat distortion tem-
mmt layers stabilize in a crystalline phase different from perature, and O2 gas permeability. These concurrent prop-
that found in the neat nylon-6, with a higher HDT.41 erty improvements are well beyond what can be generally

3.6. Oxygen Gas Permeability

Nanoclays are believed to increase the gas barrier proper-


ties by creating a maze or “tortuous path” that retards the
progress of gas molecules through the matrix resin.3, 42, 43
The mechanism for the improvement is attributed to the
increase in the tortuosity of the diffusive path for a pene-
trating molecule. Indeed, a simple tortuosity-based model
has been found to explain experimental trends satisfacto-
rily. The gas permeability of nanocomposites (PPLACN) is
related to the permeability of the pure PLA (PPLA) and the
volume fraction (fclay) and width (Dclay) of the dispersed
clay as44
Fig. 8. Oxygen gas permeability of pure PLA and various PLACN films
PPLACN 1 as a function of organoclay (C2C218-mmt) content (wt%) measured at 20 °C

PPLA 1  ( Lclay / 2 Dclay )f clay and 90% relative humidity.

7
Ray et al./Biodegradable Polylactide/Montmorillonite J. Nanosci. Nanotech. 2003, 3, 1–5

achieved through the micro/macro-composite preparation 22. From Kanebo Ltd., Japan, wave site www.kanebotx.com (accessed
or chemical modification of pure PLA. Aug. 27, 2002).
23. J. D. Gu, M. Gada, G. Kharas, D. Eberiel, S. P. McCarthy, and R. A.
Gross, Polym. Mater. Sci. Eng. 67, 351 (1992).
Acknowledgment: Thanks are due to the Japan Society for 24. N. Ogata, G. Jimenez, H. Kawai, and T. Ogihara. J. Polym. Sci. Part
B: Polym. Phys. 35, 389 (1997).
the Promotion of Science (JSPS) for the award of a postdoc-
25. S. Sinha Ray, K. Okamoto, K. Yamada, and M. Okamoto, NanoLetters
toral fellowship and a research grant to S.S.R. (no. P02152). 2, 423 (2002); S. Sinha Ray, P. Maiti, M. Okamoto, K. Yamada, and
K. Ueda, Macromolecules 35, 3104 (2002).
26. S. Sinha Ray, K. Yamada, A. Ogami, M. Okamoto, and K. Ueda,
References and Notes Macromol. Rapid Commun. 23, 943 (2002); S. Sinha Ray, K. Yamada,
M. Okamoto, A. Ogami, and K. Ueda, Chem. Mater. 15, 1456 (2003).
1. M. Biswas and S. Sinha Ray, Adv. Polym. Sci. 155, 167 (2001); 27. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, Polymer 44,
M. Alexander and P. Doubis, Mater. Sci. Eng. R28, 1 (2000). 857 (2003).
2. E. P. Giannelis, Appl. Organomet. Chem. 12, 675 (1998). 28. S. Sinha Ray, K. Yamada, M. Okamoto, Y. Fujimoto, A. Ogami, and
3. P. B. Messersmith and E. P. Giannelis, J. Polym. Sci. Part A: Polym. K. Ueda, Polymer, in press.
Chem. 33, 1047 (1995); K. Yano, A. Usuki, A. Okada, T. Kurauchi, and 29. S. Sinha Ray, M. Okamoto, K. Yamada, A. Ogami, and K. Ueda,
O. Kamogaito, J. Polym. Sci. Part A: Polym. Chem. 31, 2493 (1993). Macromolecules, in press.
[Author: 4. J. W. Gilman, Appl. Clay Sci. 15, 31 (1999). 30. We have checked the degradation of intercalated salt by using thermo-
Need 5. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, NanoLetters gravimetric (TG) analysis. Up to 210 °C, there is no degradation of
updates 2, 1093 (2002); S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, intercalated salt.
for Refs. Macromol. Mater. Eng. 288, 203 (2003). 31. P. H. Nam, P. Maiti, M. Okamoto, T. Kotaka, N. Hasegawa, and
6. R. A. Vaia, K. D. Jandt, E. J. Kramer, and E. P. Giannelis, Macro- A. Usuki, Polymer 42, 9633 (2001).
7, 28, 29,
molecules 28, 8080 (1995). 32. E. W. Fisher, H. J. Sterzel, and G. Wegner, Kolloid Z. Z. Polym. 25,
38, and 7. S. Sinha Ray and M. Okamoto, Prog. Polym. Sci., in press. 980 (1973).
42.] 8. S. W. Brindly and G. Brown, editors, Crystal Structure of Clay Min- 33. M. Okamoto, H. Kubo, and T. Kotaka, Macromolecules 31, 4223
erals and Their X-ray Diffraction, Mineralogical Society, London (1998).
(1980). 34. D  (lk)/(b cos v), where k is a constant and generally equal to 0.9,
9. P. Aranda and E. Ruiz-Hitzky, Chem. Mater. 4, 1395 (1992). l  0.154 nm, and v is the WAXD peak position. B. D. Cullity, Prin-
10. D. J. Greenland, J. Colloid. Sci. 18, 647 (1963). ciples of X-ray Diffraction, Addison-Wesley, Reading, MA (1978).
11. A. Blumstein, J. Polym. Sci. A 3, 2665 (1965). 35. Because the silicate layers are composed of heavier elements (Al, Si,
12. R. Krishnamoorti, R. A. Vaia, and E. P. Giannelis, Chem. Mater. 8, O) than the interlayer and surrounding matrix (C, H, N, etc.), they
1728 (1996). appear darker in bright-field images. R. E. Klimentidis and I. D. R.
13. J. S. Chen, M. D. Poliks, C. K. Ober, Y. Zhang, U. Wiesner, and E. P. Mackinnon, Clays Clay Miner. 34, 155 (1986).
RESEARCH ARTICLE

Giannelis, Polymer 43, 4895 (2002). 36. B. Wunderlich, Macromolecular Physics, Academic Press, New York
14. S. Sinha Ray, K. Okamoto, and M. Okamoto, Macromolecules 36, (1976), Vol. 2, p. 363.
2355 (2003). 37. P. Maiti, P. H. Nam, M. Okamoto, A. Usuki, and N. Hasegawa, Macro-
15. R. E. Drumright, P. R. Gruber, and D. E. Henton, Adv. Mater. 12, molecules 35, 2042 (2002), and references cited therein.
1841 (2000). 38. J. Y. Nam, S. Sinha Ray, and M. Okamoto, Macromolecules, in press.
16. J. Lunt, Polym. Degrad. Stab. 59, 145 (1998). 39. S. Sinha Ray, K. Okamoto, P. Maiti, and M. Okamoto, J. Nanosci.
17. S. H. Kim, Y. K. Han, Y. H. Kim, and S. I. Hong, Macromol. Chem. Nanotechnol. 2, 171 (2002).
193, 1623 (1991); H. R. Kricheldorf and A. Serra, Polym. Bull. 14, 40. A. Usuki, M. Kawasumi, Y. Kojima, A. Okada, T. Kurauchi, and
497 (1985). O. Kamigaito, J. Mater. Res. 8, 1185 (1993).
18. H. R. Kricheldorf, M. Berl, and N. Scharngal, Macromolecules 21, 41. E. Manias, A. Touny, L. Wu, K. Strawhecker, B. Lu, and T. C.
286 (1988). Chung, Chem. Mater. 13, 3516 (2001).
19. A. J. Nijenhuis, D. W. Grijpma, and A. J. Pennings, Macromolecules 42. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, Comp. Inter-
25, 6419 (1992). faces, in press.
20. O. Martin and L. Averous, Polymer 42, 6209 (2001); H. Tsuji and 43. G. W. Beall, in Polymer-Clay Nanocomposites, edited by T. J.
Y. Ikada, J. Appl. Polym. Sci. 67, 405 (1998). Pinnavaia and G. W. Beall, Wiley, New York, (2001), p. 267.
21. Q. Fang and M. A. Hanna, Ind. Crops Prod. 10, 47 (1999). 44. L. Nielsen, J. Macromol. Sci. Chem. A1(5), 929 (1967).

Received: 11 April 2003. Revised/Accepted: 1 May 2003.

You might also like